You are on page 1of 12

pathogens

Review
Listeria monocytogenes Virulence, Antimicrobial
Resistance and Environmental Persistence: A Review
Lavious Tapiwa Matereke 1,2, * and Anthony Ifeanyi Okoh 1,2, *
1 SAMRC Microbial Water Quality Monitoring Centre, University of Fort Hare, Alice 5700, South Africa
2 Applied and Environmental Microbiology Research Group, Department of Biochemistry and Microbiology,
University of Fort Hare, Alice 5700, South Africa
* Correspondence: 200909597@ufh.ac.za (L.T.M.); AOkoh@ufh.ac.za (A.I.O.)

Received: 5 May 2020; Accepted: 20 June 2020; Published: 30 June 2020 

Abstract: Listeria monocytogenes is a ubiquitous opportunistic pathogen responsible for the well-known
listeriosis disease. This bacterium has become a common contaminant of food, threatening the food
processing industry. Once consumed, the pathogen is capable of traversing epithelial barriers, cellular
invasion, and intracellular replication through the modulation of virulence factors such as internalins
and haemolysins. Mobile genetic elements (plasmids and transposons) and other sophisticated
mechanisms are thought to contribute to the increasing antimicrobial resistance of L. monocytogenes.
The environmental persistence of the pathogen is aided by its ability to withstand environmental
stresses such as acidity, cold stress, osmotic stress, and oxidative stress. This review seeks to give an
insight into L. monocytogenes biology, with emphasis on its virulence factors, antimicrobial resistance,
and adaptations to environmental stresses.

Keywords: Listeria monocytogenes; virulence; biofilm; antimicrobial resistance; environmental stress

1. Introduction
Listeria species are ubiquitous Gram-positive rods found in different environmental niches.
Among the species in the genus Listeria, Listeria monocytogenes and Listeria ivanovii, are pathogenic,
with the former being a human foodborne pathogen causing listeriosis and associated with other
human ailments such as bacteremia, encephalitis, and sepsis [1]. L. monocytogenes was proven as the
causative agent of listeriosis following an outbreak in 1989, but the pathogen had been detected as early
as 1924 [2]. It is a psychrotolerant pathogen, capable of growing at different temperatures (1–45 ◦ C),
but the optimum temperature ranges from 30 to 37 ◦ C [3]. The pathogen is proficient in switching
between saprophytism and virulence, depending on the environmental setting [4].
Infection by L. monocytogenes results in either non-invasive gastrointestinal listeriosis among
immunocompetent individuals or invasive listeriosis among immunocompromised people [5,6].
Invasive listeriosis leads to abortion in pregnant women and meningitis in immunocompromised
individuals [7]. During the 2017–2018 listeriosis outbreak in South Africa, a 42% fatality rate was
recorded, with more than 1000 laboratory-confirmed cases, and the most infected were infants and
pregnant women [8]. Individuals on medications that minimize gastric acidity as well as chronic renal
failure and cirrhosis patients are also at risk [9]. Also, Listeria-related urinary tract infections were
reported in an incident wherein L. monocytogenes was detected in urine samples [10].
The major transmission route to humans is contaminated food, mainly ready-to-eat foods [11].
Irrigation waters and agricultural soils also harbor multidrug-resistant L. monocytogenes that is likely
to be disseminated to agricultural fresh produce, posing a threat to food safety [12]. Poor hygiene
practices and inadequate implementation of standard sanitation operating procedures (SSOPs) in the
food processing industry have led to listeriosis outbreaks [13]. Furthermore, the pathogen easily forms

Pathogens 2020, 9, 528; doi:10.3390/pathogens9070528 www.mdpi.com/journal/pathogens


Pathogens 2020, 9, 528 2 of 12

biofilms and persister cells on surfaces, which may not be easily removed by standard sanitation
protocols [14]. L. monocytogenes can be found concealed in difficult-to-clean tools such as slicers,
food transport vehicles, and refrigeration units [14]. The pathogen also inhabits the gastrointestinal
tract of humans and animals; hence, unhygienic practices may facilitate its dispersal through the
contamination of food and equipment [2].
Adaptations of the pathogen to environmental stresses and pre-exposure adaptation to
sublethal concentrations of antimicrobial agents have subsequently contributed to its antimicrobial
resistance [15,16]. Therefore, understanding the principal mechanisms governing L. monocytogenes
survival, virulence, antimicrobial resistance, and persistence in adverse environmental conditions is
vital for the management and control of this pathogen, as well as the development of novel antimicrobial
agents against it.

2. Virulence Factors

2.1. Hemolysins
Production of hemolysins by Listeria species was first demonstrated by Harvey and Faber in 1941
using L. monocytogenes [17]. It was later shown that this hemolysin is an ortholog of streptolysin O
(SLO) from Streptococcus pyogenes and hence was named listeriolysin O (LLO) [18]. Encoded by the hly
gene, listeriolysin O is a pH-controlled toxin that has been shown to promote infection from several
niches within the host, mainly the extracellular environment, phagosome, and cytosol [19]. In the
extracellular environment, LLO facilitates the internalization of L. monocytogenes into phagosomes
by creating membrane perforations into the host cell [20]. LLO induces ion-permeable pores on the
plasma membrane which allow calcium influx into the host cells, and this promotes the invasion of
Hep-2 epithelial cells by L. monocytogenes [21]. LLO also mediates apoptosis of T cells, lymphocytes,
and dendritic cells [22]. Within the phagosome, LLO induces pore formation to facilitate bacterial escape
from host phagosomes (lysis) and aids the pathogen in intracellular replication [23]. The membrane
lesions created during the process enable the passage of phospholipases which hydrolyze membrane
phospholipids, resulting in complete breakdown of the plasma membrane [17]. LLO also suppresses
reactive oxygen species (ROS) released by the phagocyte in response to infection, but the mechanism is
yet to be understood [24]. In the cytosol, LLO prompts host histone modifications by deacetylation
of H4 and dephosphorylation of Ser10 in H3 histones [25]. It also enhances infection by promoting
the degradation of host proteins, though the mechanisms are yet to be understood [19]. In one study,
treatment of the host proteome with LLO reduced the abundance of 149 proteins [26]. LLO causes
mitochondrial fragmentation by inducing calcium influx through the ion-permeable pores on the
plasma membrane [27]. However, overexpression of LLO can expose the pathogen to host immune
defenses. This is a result of membrane lesions which lead to apoptosis and demolition of the
pathogen’s intracellular niche, thus exposing the pathogen to the circulating defense machinery [28,29].
Host proteasomal degradation of LLO can be one mechanism employed to limit its activity within the
host cytosol [28]. Another study also revealed in vitro aggregation and denaturation of LLO at 37 ◦ C
and neutral pH, suggesting that this could be another regulatory mechanism for LLO activity within
the host cytosol [30].

2.2. Phospholipases
L. monocytogenes secretes phosphatidylinositol-specific phospholipase C (PI-PLC or PLC-A)
encoded by the plcA gene and non-specific phosphotidylcholine phospholipase C (PC-PLC or PLC-B)
encoded by the plcB gene [31]. PLC-A (PI-PLC) aids the pathogen in exiting the phagosome into
the cytosol [32]. Together with PC-PLC and LLO, PI-PLC also facilitates the pathogen’s exit from
the double-membrane vacuole formed upon cell-to-cell spread [33]. Both PI-PLC and PC-PLC aid
L. monocytogenes in subverting autophagy-mediated clearance by inhibiting pre-autophagosomal
maturation or target recognition by the degradative pathway [34]. Autophagy is the process through
Pathogens 2020, 9, 528 3 of 12

which host intracellular dysfunctional organelles are degraded and recycled [35]. Autophagy also
targets invading pathogens and thus plays a role in innate immunity [35]. In one study, mutant
L. monocytogenes strains incapable of expressing phospholipases were vulnerable to autophagy during
macrophage infection [36].
PC-PLC is initially expressed as a zymogen which is then activated by a zinc metalloprotease under
acidic environments [37]. It assists in vacuolar escape in times of LLO deficiency during its invasion of
epithelial cells, and its activity is crucial for cell-to-cell spread [38]. PC-PLC dual enzymatic activity as
a phospholipase and sphingomyelinase could be essential for L. monocytogenes dispersal inside the
host [39]. PC-PLC is an essential pathogenicity factor contributing to L. monocytogenes murine cerebral
listeriosis [40]. However, premature discharge of PC-PLC into the host cell cytosol can be lethal to
L. monocytogenes; thus, the pathogen has a PC-PLC regulation mechanism for efficient pathogenicity [41].
A PC-PLC mutant exhibited low resistance against intracellular killing by neutrophils, suggesting
PC-PLC may potentiate the action of neutrophils against L. monocytogenes [42]. L. monocytogenes
regulates the activity of PC-PLC by increasing the pH within the double-membrane vacuole [41].
This inhibits the activity of the metalloprotease, resulting in an inactive PC-PLC [43].

2.3. ActA
Encoded by the actA gene, the protein actA enables actin recruitment and polymerization,
resulting in actin-based intracellular motility in pathogenic Listeria species [17]. Polymerized actin
close to the phagosome membrane possibly gains entry through listeriolysin-mediated pores, and its
polymerization could widen the pores [44]. This results in phagosome disruption, promoting bacterial
escape from the phagosome [44]. Actin propels L. monocytogenes to host cell membranes, resulting in
elongated protrusions from the membrane (fibroids) encircling the bacteria and extending to adjacent
cells which engulf the fibroids [17]. In epithelial cells, ActA camouflages the pathogen with host
proteins, thus shielding it from autophagy [45]. Using a murine model, one study demonstrated that
ActA is crucial for placental invasion by L. monocytogenes and facilitates its vertical dissemination to
the fetus [46]. ActA facilitates aggregation in L. monocytogenes through uninterrupted ActA–ActA
interactions, is involved in biofilm formation, and mediates intestinal colonization [47]. Results from
one study demonstrated that actin-based intracellular motility facilitates the exit of the pathogen from
autophagic membranes within the macrophage cytosol [48].

2.4. Internalins
Internalins are surface proteins encoded by genes which are associated with virulence in pathogenic
Listeria species. Internalin A (InlA) and Internalin B (InlB), both encoded by the inlAB operon, were the
first internalins to be characterized from L. monocytogenes [17]. E-cadherin and Met are the internalins’
receptors, respectively, both located on host cell surfaces [49]. Internalins mediate the adhesion and
internalization of the pathogen by host nonphagocytic cells [50]. InlA-mediated internalization is
constrained to a limited population of epithelial cells, and InlB-mediated internalization targets various
cells like hepatocytes and epithelial and endothelial cells [51]. Using an epidemic strain, one study
demonstrated that InlB enhances infection of the liver and spleen by L. monocytogenes [52].
Internalins consist of a signal peptide at the N-terminus and 22-amino acid leucin-rich-repeats
(LRR) [53]. The N-terminal region alone is capable of enhancing bacterial penetration into permissive
cells [54]. These LLR regions are responsible for protein–protein interactions and are present in both
prokaryotic and eukaryotic cells, where they also facilitate adhesion [55]. Other pathogenic bacterial
species also express LLR regions [50]. Interactions between Internalin A and E-cadherin are vital
for traversing the placental and intestinal barriers [49]. The novel InlL facilitates L. monocytogenes
attachment to abiotic surfaces and plays a role in biofilm establishment [56]. Internalin K recruits a
major vault protein (MVP) that camouflages the pathogen, shielding it from autophagic recognition [57].
Other novel internalins (InlP1, InlPq, and InlP4) which contribute to the virulence of L. monocytogenes,
have been reported [58].
Pathogens 2020, 9, 528 4 of 12

3. Antimicrobial Resistance

3.1. Horizontal Gene Transfer


Listeria species are reportedly capable of exchanging resistance determinants with other bacterial
species. The broad-host-range Streptococcus agalactiae plasmid IP501, which confers resistance against
macrolides, lincosamides, chloramphenicol, and streptogramins, was reportedly transferrable to
L. monocytogenes via conjugation [59,60]. The same plasmid was reportedly capable of replication within
Listeria species and transference among different Listeria species as well as back to Streptococcus [61].
The antimicrobial resistance plasmid pAM Beta I (pAMβ1) was successfully transferred from
Streptococcus faecalis to some L. monocytogenes isolates by conjugation [62]. In the same study,
conjugates were found to replicate within the host and were transferrable among L. monocytogenes
strains or back to Streptococcus faecalis. One study demonstrated conjugation of the tet(S) gene from
Lactococcus garvieae, a fish pathogen, to L. monocytogenes [63]. The transposon Tn916 harboring the
tet(M) gene, originally discovered in Enterococcus faecalis, was demonstrated to be transferrable
between Enterococcus faecalis and Listeria innocua and that it was capable of jumping from one host
to the other among various Listeria species [64]. Another Tn916-like transposon, TnFO1, harbored by
multidrug-resistant Enterococcus faecalis, was found to be transferred from Enterococcus. faecalis to some
species, including Listeria innocua, via conjugation [65].
Plasmids and transposons are thought to mediate tetracycline resistance in Listeria species [66].
TetM genes discovered in two tetracycline-resistant L. monocytogenes isolates were analogous to the
ones previously discovered in Staphylococcus aureus [67]. In the same study, TetM genes from the other
two isolates had sequences that were closely related to those of a family of conjugative transposons
(Tn916-Tn1545). The Tn916-Tn1545 family is closely related to SHGII, which harbors TetM genes
in Enterococci.

3.2. Persister Cells and Biofilm Formation


One mechanism underlying the antimicrobial resistance in L. monocytogenes is the formation of
persister cells and biofilms. Persisters are a fraction of the overall cell population that is in a dormant,
non-dividing state, which enhances them to withstand bactericidal antibiotics and harsh environmental
conditions [68]. Persister cells provide an endurance strategy to L. monocytogenes, which is transformed
from bacilli to cocci during the transition [69]. The persister phenomenon probably contributes to
L. monocytogenes survival against disinfectants and may shield the pathogen from the host defense
system for longer periods, which explains the absence of listeriosis clinical symptoms several days after
infection [68]. In one study, long-term-survival cells were extremely resistant to high temperatures
and pressure stresses [69]. These persisters are also able to withstand preservatives used to control
heat-resistant bacteria during food processing [70].
A biofilm is a population of bacterial cells attached to each other or a surface and enclosed
by a self-produced matrix of extracellular polymeric substances (EPS) (proteins, polysaccharides,
and extracellular DNA) [71]. The matrix is a nutrient reservoir, facilitates adhesion, acts as a
protective barrier against heat and desiccation, heavy metals, ultraviolet rays, acids, and also inhibits
the permeation of antimicrobial agents by restricting their diffusion potentials [72,73]. Persistent
L. monocytogenes strains were found to have significantly better biofilm-forming capabilities than
non-persistent ones [74]. Biofilms are reportedly behind the pathogen’s increased tolerance to
quaternary ammonium compounds due to changes in membrane fluidity [75]. Biofilms can also
protect L. monocytogenes from the action of biocides, but the underlying mechanism of biocide
protection is yet to be clarified [76].

3.3. Efflux Pumps


The minimum susceptibility of L. monocytogenes to antimicrobial agents is also attributed to some
efflux pumps. The Lde efflux pump was found to be associated with fluoroquinolones resistance in
Pathogens 2020, 9, 528 5 of 12

L. monocytogenes isolated from listeriosis patients [77]. In the same study, the protein sequence of
the Lde pump was found to be 44% similar to that of PmrA of Streptococcus pneumoniae, a secondary
multidrug transporter of the major facilitator superfamily (MFS). Evidence from another study also
suggests that the Lde efflux pump contributes to ciprofloxacin resistance in L. monocytogenes [78]. In the
same study, minimum inhibitory concentrations (MICs) of the disinfectant benzalkonium chloride (BC)
were unaffected by the absence of the lde gene, indicating that the Lde efflux pump may not confer BC
resistance. An earlier study had reported that both the Lde and the MdrL efflux pumps may confer BC
resistance to L. monocytogenes, alongside other mechanisms [79]. Increased L. monocytogenes tolerance
to BC can be a result of the expression of an EmrE efflux pump [80]. The MdrL efflux pump is in charge
of the extrusion of antimicrobial agents and heavy metals from the pathogen [81]. The MdrM and
MdrT efflux pumps enable persistence and replication of the pathogen within the gastrointestinal tract,
counteracting the bactericidal effects of mammalian bile [82]. The bile component cholic acid induces
the expression of both pumps, but MdrT regulates the extrusion of cholic acid, which can be lethal to
mutant strains lacking the MdrT efflux pump [83].

3.4. Environmental Stress Exposure


Listeria species may encounter sublethal concentrations of antimicrobial agents and growth
inhibitors, e.g., bacteriocins and disinfectants (biocides) in the food processing and agriculture
industries [84]. Such exposure may trigger adaptation to resist higher concentrations of these
antimicrobials. Pre-exposure adaptation subsequently contributes to antibiotic resistance of Listeria
species [15]. Co-selection for antibiotic resistance of Listeria species resulting from the use of biocides
and heavy metals is reportedly a general concern [85]. One study reported significant correlations
between biocide tolerance and antibiotic resistance in Listeria among other bacterial species isolated
from seafood [86]. Unlimited exposure of the pathogen to sublethal dosages of chlorine may indirectly
induce a significant tolerance to antibiotics, due to alterations in cell structure and morphology [87].
The bacteriocin nisin produced by Lactococcus lactis is one of the antimicrobial agents commonly applied
in the food processing industry. However, nisin resistance by L. monocytogenes has been reported [88].
Nisin targets cell wall biosynthesis and cell membrane disruption, thus resistance may arise from
alterations in the cell wall composition, hindering bacteriocin entry into the cell [89].
L. monocytogenes encounters environmental stresses in the food processing chain, and these stresses
are usually effects of the numerous food preservation methods such as high salinity and acidic pH [90].
Adaptation of the pathogen to these sublethal stressful environmental conditions has been reported to
contribute to antimicrobial resistance [91]. This is usually a result of bacterial phenotypic changes,
mainly affecting cell wall and membrane structures, or alterations in efflux pump activity [92].

4. Environmental Stress Adaptation

4.1. Acidity
Acid tolerance response (ATR), arginine deiminase (ADI), and the glutamate decarboxylase
(GAD) system are the three mechanisms employed by L. monocytogenes for persistence in low pH
environments [93,94]. The ATR system is activated by pre-exposure to sublethal pH (5–6) and facilitates
the pathogen’s endurance in highly acidic settings [95]. The two-component pathway also protects the
pathogen from osmotic stress and high temperatures [2].
The ADI system is operated by three enzymes: arginine deiminase hydrolyzes arginine to citrulline
and ammonia, ornithine carbamoyltransferase catalyzes the conversion of citrulline to ammonia and
carbamoylphospate, and carbamate kinase synthesizes ATP using carbamoylphosphate and adenosine
diphosphate (ADP) [94]. The GAD system promotes the survival and growth of the pathogen in mild
and extremely acidic environments [96]. It also enhances its passage through the stomach (gastric
environment), enabling access to and invasion of the intestinal epithelial cells. Under extremely acidic
Pathogens 2020, 9, 528 6 of 12

conditions, the enzyme glutamate decarboxylase (encoded by gadA or gadB) decarboxylates imported
extracellular glutamic acid to γ-aminobutyrate (GABA), consuming intracellular proteins [97].
Another mechanism for acid tolerance is the F0F1–ATPase complex. The F0F1–ATPase enzyme
has two different domains: a cytoplasmic catalytic portion (F1 ) responsible for ATP synthesis or
hydrolysis and the integral membrane domain (F0 ), functioning as a proton channel [98]. The enzyme
synthesizes ATP aerobically using protons passing into the cell cytoplasm (oxidative phosphorylation)
or hydrolyzes ATP as protons exit the cell, generating a proton motive force (PMF) [99].

4.2. Osmotic Stress


L. monocytogenes is reportedly capable of withstanding high salt concentrations (NaCl) [100].
In response to osmotic stress, the pathogen activates two different mechanisms: the primary and
the secondary responses. During the primary response, there is an influx of potassium cations
and glutamate into the cell [89]. Uptake of osmoprotectants such as glycine betaine and carnitine
follows [101]. Both mechanisms also assist the bacterium in maintaining turgidity and help in the
stabilization of protein structure and function [102]. Two genes (lmo1078 and lmo2085) are responsible
for cell envelop modification as an adaptation to osmotic stress [103,104]. A glucose phosphorylase,
lmo1078, synthesizes UDP-glucose, the building unit of membrane glycolipids and the cell wall [103].
lmo2085 is an exterior protein covalently bound to cell wall peptidoglycans [104]. General stress
response proteins are also expressed in response to osmotic shock, and in the absence of osmoprotectants,
the protein Ctc, which confers high osmolarity resistance, is expressed [105].

4.3. Low Temperatures (Cold Stress)


Adaptation of L. monocytogenes to low temperatures involves the expression of cold shock proteins
(CSPs) [106]. These are small proteins which assume the role of molecular chaperones, enabling
replication, transcription, translation, and protein folding at low temperatures [107]. Another adaptation
mechanism is the importation of glycine betaine and carnitine as cryoprotectants [90]. Increased
intracellular solutes levels help reduce the loss of intracellular water in times of low temperatures [89].
In one investigation, a mutant L. monocytogenes strain for glycine betaine and carnitine uptake systems
failed to accumulate or transport any of the two and, thus, it could not survive at low temperatures [108].
L. monocytogenes also alter its membrane structure at low temperatures. The membrane’s rigidity
increases due to an upsurge in the concentration of unsaturated fatty acids [109]. This prevents the
development of a gel-like structure which may allow leakage of the cytoplasmic content [110].

4.4. Oxidative Stress


Oxidative stress arises from exposure to ROS (superoxide, hydroxyl radicals, and hydrogen
peroxide) which may be generated during food processing [111]. Adaptations to oxidative stress
includes ROS detoxification systems such as catalase (Cat), superoxide dismutase (Sod), and alkyl
hydroperoxidase (AhpCF) [112]. Ferritin-like protein (fri) reportedly protects L. monocytogenes from the
oxidative effects of hydrogen peroxide [113].

5. Conclusions
L. monocytogenes undoubtedly remains a foodborne pathogen of public health concern, owing to the
high mortality and hospitalization rates caused by its infection [73]. The 2017–2018 listeriosis outbreak
in South Africa was a clear evidence that, despite the development of treatments and biocontrol agents
to combat L. monocytogenes, the pathogen remains a threat to food safety [114]. This is further worsened
by the increasing tolerance of the pathogen to biocontrol agents and antimicrobials. Environmental
stresses also have an impact on the pathogen’s physiology, morphology, pathogenicity, gene expression,
and antimicrobial resistance [91]. Understanding the pathogen’s virulence, antimicrobial resistance
mechanisms, and environmental stress adaptation will significantly contribute to the development of
novel efficient, cost-effective antimicrobial agents and biocontrol methods for combating the pathogen.
Pathogens 2020, 9, 528 7 of 12

Continuous monitoring, stringent surveillance, and source tracking are also crucial for the detection of
the pathogen’s antimicrobial resistance developments [15].

Author Contributions: L.T.M. conducted the literature search and prepared the original manuscript. A.I.O.
provided the conceptualization, reviewing, editing, supervision, and sourced the funding. All authors have read
and agreed to the published version of the manuscript.
Funding: This research was funded by the South African Medical Research Council (Grant number: SAMRC/UFH/P790).
Acknowledgments: The authors are grateful to the South African Medical Research Council for financial support.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Volokhov, D.; Rasooly, A.; Chumakov, K.; Chizhikov, V. Identification of Listeria Species by Microarray-Based
Assay. J. Clin. Microbiol. 2002, 40, 4720–4728. [CrossRef]
2. Davis, M.L.; Ricke, S.C.; Donaldson, J.R. Establishment of Listeria monocytogenes in the Gastrointestinal Tract.
Microorganisms 2019, 7, 75. [CrossRef] [PubMed]
3. Pereira, S.A.; Alves, A.; Ferreira, V.; Teixeira, C.M. The Impact of Environmental Stresses in the Virulence
Traits of Listeria monocytogenes Relevant to Food Safety. In Listeria monocytogenes; IntechOpen: Rijeka, Croatia,
2018; pp. 1–21.
4. David, D.J.V.; Cossart, P. Recent advances in understanding Listeria monocytogenes infection: The importance
of subcellular and physiological context. F1000Research 2017, 6, F1000 Faculty Rev-1126. [CrossRef] [PubMed]
5. Allen, K.J.; Wałecka-Zacharska, E.; Chen, J.C.; Katarzyna, K.-P.; Devlieghere, F.; Van Meervenne, E.; Osek, J.;
Wieczorek, K.; Bania, J. Listeria monocytogenes—An examination of food chain factors potentially contributing
to antimicrobial resistance. Food Microbiol. 2016, 54, 178–189. [CrossRef]
6. Iwu, C.D.; Okoh, A.I. Characterization of antibiogram fingerprints in Listeria monocytogenes recovered from
irrigation water and agricultural soil samples. PLoS ONE 2020, 15, e0228956. [CrossRef]
7. Cossart, P.; Archambaud, C. The bacterial pathogen Listeria monocytogenes: An emerging model in prokaryotic
transcriptomics. J. Biol. 2009, 8, 107. [CrossRef]
8. Desai, A.N.; Anyoha, A.; Madoff, L.C.; Lassmann, B. Changing epidemiology of Listeria monocytogenes
outbreaks, sporadic cases, and recalls globally: A review of ProMED reports from 1996 to 2018. Int. J.
Infect. Dis. 2019, 84, 48–53. [CrossRef]
9. Ramaswamy, V.; Cresence, V.M.; Rejitha, J.S.; Lekshmi, M.U.; Dharsana, K.S.; Prasad, S.P.; Vijila, H.M.
Listeria—Review of epidemiology and pathogenesis. J. Microbiol. Immunol. Infect. 2007, 40, 4–13.
10. Danion, F.; Maury, M.M.; Leclercq, A.; Moura, A.; Perronne, V.; Léotard, S.; Dary, M.; Tanguy, B.;
Bracq-Dieye, H.; Thouvenot, P.; et al. Listeria monocytogenes isolation from urine: A series of 15 cases
and review. Clin. Microbiol. Infect. 2017, 23, 583–585. [CrossRef]
11. Kurpas, M.; Wieczorek, K.; Osek, J. Ready-to-eat Meat Products as a Source of Listeria monocytogenes.
J. Vet. Res. 2018, 62, 49–55. [CrossRef]
12. Iwu, D.C.; Okoh, I.A. Preharvest Transmission Routes of Fresh Produce Associated Bacterial Pathogens with
Outbreak Potentials: A Review. Int. J. Environ. Res. Public Health 2019, 16, 4407. [CrossRef] [PubMed]
13. Dufour, C. Application of EC regulation no. 2073/2005 regarding Listeria monocytogenes in ready-to-eat foods
in retail and catering sectors in Europe. Food Control. 2011, 22, 1491–1494. [CrossRef]
14. Lakicevic, B.; Nastasijevic, I. Listeria monocytogenes in retail establishments: Contamination routes and control
strategies. Food Rev. Int. 2017, 33, 247–269. [CrossRef]
15. Olaimat, A.N.; Al-holy, M.A.; Shahbaz, H.M.; Al-nabulsi, A.A.; Abu Ghoush, M.H.; Osaili, T.M.; Ayyash, M.M.;
Holley, R.A. Emergence of Antibiotic Resistance in Listeria monocytogenes Isolated from Food Products:
A Comprehensive Review. Compr. Rev. Food Sci. Food Saf. 2018, 17, 1277–1292. [CrossRef]
16. Lungu, B.; Ricke, S.C.; Johnson, M.G. Growth, survival, proliferation and pathogenesis of Listeria monocytogenes
under low oxygen or anaerobic conditions: A review. Anaerobe 2009, 15, 7–17. [CrossRef] [PubMed]
17. Va’Zquez-Boland, J.; Kuhn, M.; Berche, P.; Chakraborty, T.; Domi’Nguez-Bernal, G.; Goebel, W.;
Gonzalez-Zorn, B.; Wehland, J.; Kreft, J. Listeria Pathogenesis and Molecular Virulence Determinants.
Clin. Microbiol. Rev. 2001, 14, 584–640. [CrossRef]
Pathogens 2020, 9, 528 8 of 12

18. Farber, J.M.; Peterkin, P.I. Listeria monocytogenes, a Food-Borne Pathogen. Mcrobiol. Rev. 1991, 55, 476–511.
[CrossRef]
19. Osborne, S.E.; Brumell, J.H. Listeriolysin O: From bazooka to Swiss army knife. Philos. Trans. R Soc. B
Biol. Sci. 2017, 372, 20160222. [CrossRef]
20. Vadia, S.; Arnett, E.; Haghighat, A.-C.; Wilson-Kubalek, E.M.; Tweten, R.K.; Seveau, S. The pore-forming toxin
listeriolysin O mediates a novel entry pathway of L. monocytogenes into human hepatocytes. PLoS Pathog.
2011, 7, e1002356. [CrossRef]
21. Dramsi, S.; Cossart, P. Listeriolysin O-Mediated Calcium Influx Potentiates Entry of Listeria monocytogenes
into the Human Hep-2 Epithelial Cell Line. Infect. Immun. 2003, 71, 3614–3618. [CrossRef]
22. Carrero, J.A.; Calderon, B.; Unanue, E.R. Listeriolysin O from Listeria monocytogenes Is a Lymphocyte
Apoptogenic Molecule. J. Immunol. 2004, 172, 4866–4874. [CrossRef] [PubMed]
23. Maury, M.M.; Chenal-Francisque, V.; Bracq-Dieye, H.; Han, L.; Leclercq, A.; Vales, G.; Moura, A.; Gouin, E.;
Scortti, M.; Disson, O.; et al. Spontaneous Loss of Virulence in Natural Populations of Listeria monocytogenes.
Infect. Immun. 2017, 85, e00541-17. [CrossRef] [PubMed]
24. Lam, G.Y.; Fattouh, R.; Muise, A.M.; Grinstein, S.; Higgins, D.E.; Brumell, J.H. Listeriolysin O Suppresses
Phospholipase C-Mediated Activation of the Microbicidal NADPH Oxidase to Promote Listeria monocytogenes
Infection. Cell Host Microbe 2011, 10, 627–634. [CrossRef] [PubMed]
25. Hamon, M.A.; Batsché, E.; Régnault, B.; Tham, T.N.; Seveau, S.; Muchardt, C.; Cossart, P. Histone modifications
induced by a family of bacterial toxins. Proc. Natl. Acad. Sci. USA 2007, 104, 13467–13472. [CrossRef]
[PubMed]
26. Malet, J.K.; Impens, F.; Carvalho, F.; Hamon, M.A.; Cossart, P.; Ribet, D. Rapid Remodeling of the Host
Epithelial Cell Proteome by the Listeriolysin O (LLO) Pore-forming Toxin. Mol. Cell. Proteom. 2018, 17,
1627–1636. [CrossRef]
27. Stavru, F.; Bouillaud, F.; Sartori, A.; Ricquier, D.; Cossart, P. Listeria monocytogenes transiently alters
mitochondrial dynamics during infection. Proc. Natl. Acad. Sci. USA 2011, 108, 3612–3617. [CrossRef]
28. Schnupf, P.; Portnoy, D.A. Listeriolysin O: A phagosome-specific lysin. Microbes Infect. 2007, 9, 1176–1187.
[CrossRef]
29. Hamon, M.A.; Ribet, D.; Stavru, F.; Cossart, P. Listeriolysin O: The Swiss army knife of Listeria. Trends Microbiol.
2012, 20, 360–368. [CrossRef]
30. Schuerch, D.W.; Wilson-Kubalek, E.M.; Tweten, R.K. Molecular basis of listeriolysin O pH dependence.
Proc. Natl. Acad. Sci. USA 2005, 102, 12537–12542. [CrossRef]
31. Smith, G.A.; Marquis, H.; Jones, S.; Johnston, N.C.; Portnoy, D.A.; Goldfine, H. The two distinct phospholipases
C of Listeria monocytogenes have overlapping roles in escape from a vacuole and cell-to-cell spread.
Infect. Immun. 1995, 63, 4231–4237. [CrossRef]
32. Camilli, A.; Tilney, L.G.; Portnoy, D.A. Dual roles of plcA in Listeria monocytogenes pathogenesis. Mol. Microbiol.
1993, 8, 143–157. [CrossRef] [PubMed]
33. Alberti-Segui, C.; Goeden, K.R.; Higgins, D.E. Differential function of Listeria monocytogenes listeriolysin O
and phospholipases C in vacuolar dissolution following cell-to-cell spread. Cell. Microbiol. 2007, 9, 179–195.
[CrossRef] [PubMed]
34. Tattoli, I.; Sorbara, M.T.; Yang, C.; Tooze, S.A.; Philpott, D.J.; Girardin, S.E. Listeria phospholipases subvert
host autophagic defenses by stalling pre-autophagosomal structures. EMBO J. 2013, 32, 3066–3078. [CrossRef]
35. Tattoli, I.; Sorbara, M.T.; Philpott, D.J.; Girardin, S.E. Bacterial autophagy: The trigger, the target and the
timing. Autophagy 2012, 8, 1848–1850. [CrossRef] [PubMed]
36. Birmingham, C.L.; Canadien, V.; Gouin, E.; Troy, E.B.; Cossart, P.; Higgins, D.E.; Brumell, J.H.
Listeria monocytogenes Evades Killing by Autophagy during Colonization of Host Cells. Autophagy 2007, 3,
442–451. [CrossRef] [PubMed]
37. Yeung, P.S.M.; Zagorski, N.; Marquis, H. The Metalloprotease of Listeria monocytogenes Controls Cell Wall
Translocation of the Broad-Range Phospholipase C. J. Bacteriol. 2005, 187, 2601–2608. [CrossRef] [PubMed]
38. Gründling, A.; Gonzalez, M.D.; Higgins, D.E. Requirement of the Listeria monocytogenes broad-range
phospholipase PC-PLC during infection of human epithelial cells. J. Bacteriol. 2003, 185, 6295–6307. [CrossRef]
39. Montes, L.-R.; Goñi, F.M.; Johnston, N.C.; Goldfine, H.; Alonso, A. Membrane Fusion Induced by the
Catalytic Activity of a Phospholipase C/Sphingomyelinase from Listeria monocytogenes. Biochemistry 2004, 43,
3688–3695. [CrossRef]
Pathogens 2020, 9, 528 9 of 12

40. Schlüter, D.; Domann, E.; Buck, C.; Hain, T.; Hof, H.; Chakraborty, T.; Deckert-Schlüter, M. Phosphatidylcholine-
specific phospholipase C from Listeria monocytogenes is an important virulence factor in murine cerebral
listeriosis. Infect. Immun. 1998, 66, 5930–5938. [CrossRef]
41. Marquis, H.; Hager, E.J. pH-regulated activation and release of a bacteria-associated phospholipase C during
intracellular infection by Listeria monocytogenes. Mol. Microbiol. 2000, 35, 289–298. [CrossRef]
42. Blank, B.S.; Abi Abdallah, D.S.; Park, J.J.; Nazarova, E.V.; Pavinski Bitar, A.; Maurer, K.J.; Marquis, H.
Misregulation of the broad-range phospholipase C activity increases the susceptibility of Listeria monocytogenes
to intracellular killing by neutrophils. Microbes Infect. 2014, 16, 104–113. [CrossRef] [PubMed]
43. Yeung, P.S.M.; Na, Y.; Kreuder, A.J.; Marquis, H. Compartmentalization of the Broad-Range Phospholipase C
Activity to the Spreading Vacuole Is Critical for Listeria monocytogenes Virulence. Infect. Immun. 2007, 75,
44–51. [CrossRef] [PubMed]
44. Poussin, M.A.; Goldfine, H. Evidence for the involvement of ActA in maturation of the Listeria monocytogenes
phagosome. Cell Res. 2009, 20, 109–112. [CrossRef] [PubMed]
45. Yoshikawa, Y.; Ogawa, M.; Hain, T.; Chakraborty, T.; Sasakawa, C. Listeria monocytogenes ActA is a key player
in evading autophagic recognition. Autophagy 2009, 5, 1220–1221. [CrossRef] [PubMed]
46. Le Monnier, A.; Autret, N.; Join-lambert, O.F.; Jaubert, F.; Charbit, A.; Berche, P.; Kayal, S. ActA Is Required for
Crossing of the Fetoplacental Barrier by Listeria monocytogenes. Infect. Immun. 2007, 75, 950–957. [CrossRef]
47. Travier, L.; Guadagnini, S.; Gouin, E.; Dufour, A.; Chenal-Francisque, V.; Cossart, P.; Olivo-Marin, J.; Ghigo, J.;
Disson, O.; Lecuit, M. ActA Promotes Listeria monocytogenes Aggregation, Intestinal Colonization and
Carriage. PLoS Pathog. 2013, 9, e1003131. [CrossRef]
48. Cheng, M.I.; Chen, C.; Engström, P.; Portnoy, D.A.; Mitchell, G. Actin-based motility allows Listeria
monocytogenes to avoid autophagy in the macrophage cytosol. Cell. Microbiol. 2018, 20, e12854. [CrossRef]
49. Bonazzi, M.; Lecuit, M.; Cossart, P. Listeria monocytogenes internalin and E-cadherin: From bench to bedside.
Cold Spring Harb. Perspect. Biol. 2009, 1, a003087. [CrossRef]
50. Bierne, H.; Sabet, C.; Personnic, N.; Cossart, P. Internalins: A complex family of leucine-rich repeat-containing
proteins in Listeria monocytogenes. Microbes Infect. 2007, 9, 1156–1166. [CrossRef]
51. Bierne, H.; Cossart, P. InlB, a surface protein of Listeria monocytogenes that behaves as an invasin and a growth
factor. J. Cell Sci. 2002, 115, 3357–3367.
52. Quereda, J.J.; Rodríguez-Gómez, I.M.; Meza-Torres, J.; Gómez-Laguna, J.; Nahori, M.A.; Dussurget, O.;
Carrasco, L.; Cossart, P.; Pizzaro-Cerda, J. Reassessing the role of internalin B in Listeria monocytogenes
virulence using the epidemic strain F2365. Clin. Microbiol. Infect. 2019, 25, 252.e1–252.e4. [CrossRef]
53. Seveau, S.; Pizarro-Cerda, J.; Cossart, P. Molecular mechanisms exploited by Listeria monocytogenes during
host cell invasion. Microbes Infect. 2007, 9, 1167–1175. [CrossRef]
54. Lecuit, M.; Ohayon, H.; Braun, L.; Mengaud, J.; Cossart, P. Internalin of Listeria monocytogenes with an
intact leucine-rich repeat region is sufficient to promote internalization. Infect. Immun. 1997, 65, 5309–5319.
[CrossRef]
55. Marino, M.; Braun, L.; Cossart, P.; Ghosh, P. A framework for interpreting the leucine-rich repeats of the
Listeria internalins. Proc. Natl. Acad. Sci. USA 2000, 97, 8784–8788. [CrossRef] [PubMed]
56. Popowska, M.; Krawczyk-Balska, A.; Ostrowski, R.; Desvaux, M. InlL from Listeria monocytogenes Is Involved
in Biofilm Formation and Adhesion to Mucin. Front. Microbiol. 2017, 8, 660. [CrossRef] [PubMed]
57. Loo, K.Y.; Letchumanan, V.; Dhanoa, A.; Law, W.; Pusparajah, P.; Goh, B.; Ser, H.; Wong, S.H.; Mutalib, N.A.;
Chan, K.; et al. Exploring the Pathogenesis, Clinical Characteristics and Therapeutic Regimens of
Listeria monocytogenes. Acta Sci. Microbiol. 2020, 3. [CrossRef]
58. Harter, E.; Lassnig, C.; Wagner, E.M.; Zaiser, A.; Wagner, M.; Rychli, K. The Novel Internalins InlP1 and InlP4
and the Internalin-Like Protein InlP3 Enhance the Pathogenicity of Listeria monocytogenes. Front. Microbiol.
2019, 10, 1644. [CrossRef]
59. Evans, R.P., Jr.; Macrina, F.L. Streptococcal R plasmid pIP501: Endonuclease site map, resistance determinant
location, and construction of novel derivatives. J. Bacteriol. 1983, 154, 1347–1355. [CrossRef]
60. Charpentier, E.; Courvalin, P. Antibiotic resistance in Listeria spp. Antimicrob. Agents Chemother. 1999, 43,
2103–2108. [CrossRef]
61. Vicente, M.F.; Baquero, F.; Pérez-Diaz, J.C. Conjugative acquisition and expression of antibiotic resistance
determinants in Listeria spp. J. Antimicrob. Chemother. 1988, 21, 309–318. [CrossRef]
Pathogens 2020, 9, 528 10 of 12

62. Flamm, R.K.; Hinrichs, D.J.; Thomashow, M.F. Introduction of pAM beta 1 into Listeria monocytogenes by
conjugation and homology between native L. monocytogenes plasmids. Infect. Immun. 1984, 44, 157–161.
[CrossRef] [PubMed]
63. Guglielmetti, E.; Korhonen, J.M.; Heikkinen, J.; Morelli, L.; Von Wright, A. Transfer of plasmid-mediated
resistance to tetracycline in pathogenic bacteria from fish and aquaculture environments. FEMS Microbiol. Lett.
2009, 293, 28–34. [CrossRef]
64. Celli, J.; Trieu-Cuot, P. Circularization of Tn916 is required for expression of the transposon-encoded transfer
functions: Characterization of long tetracycline-inducible transcripts reading through the attachment site.
Mol. Microbiol. 1998, 28, 103–117. [CrossRef]
65. Perreten, V.; Kollöffel, B.; Teuber, M. Conjugal Transfer of the Tn916-like Transposon TnFO1 from Enterococcus
faecalis Isolated from Cheese to Other Gram-positive Bacteria. Syst. Appl. Microbiol. 1997, 20, 27–38. [CrossRef]
66. Luque-sastre, L.; Arroyo, C.; Fox, E.M.; Mcmahon, B.J.; Bai, L.I.; Li, F.; Fanning, S. Antimicrobial Resistance
in Listeria Species. Microbiol. Spectr. 2018, 6, 237–259.
67. Bertrand, S.; Huys, G.; Yde, M.; D’Haene, K.; Tardy, F.; Vrints, M.; Swings, J.; Collard, J. Detection and
characterization of tet (M) in tetracycline-resistant Listeria strains from human and food-processing origins
in Belgium and France. J. Med Microbiol. 2005, 54, 1151–1156. [CrossRef]
68. Knudsen, G.M.; Ng, Y.; Gram, L. Survival of Bactericidal Antibiotic Treatment by a Persister Subpopulation
of Listeria monocytogenes. Appl. Environ. Microbiol. 2013, 79, 7390–7397. [CrossRef]
69. Wen, J.; Deng, X.; Li, Z.; Dudley, E.G.; Anantheswaran, R.C.; Knabel, S.J.; Zhang, W. Transcriptomic Response
of Listeria monocytogenes during the Transition to the Long-Term-Survival Phase. Appl. Environ. Microbiol.
2011, 77, 5966–5972. [CrossRef]
70. Wood, T.K.; Knabel, S.J.; Kwan, B.W. Bacterial Persister Cell Formation and Dormancy. Appl. Environ. Microbiol.
2013, 79, 7116–7121. [CrossRef]
71. Lee, B.-H.; Hébraud, M.; Bernardi, T. Increased Adhesion of Listeria monocytogenes Strains to Abiotic Surfaces
under Cold Stress. Front. Microbiol. 2017, 8, 2221. [CrossRef]
72. da Silva, E.P.; De Martinis, E.C.P. Current knowledge and perspectives on biofilm formation: The case of
Listeria monocytogenes. Appl. Microbiol. Biotechnol. 2013, 97, 957–968. [CrossRef]
73. Gray, J.A.; Chandry, P.S.; Kaur, M.; Kocharunchitt, C.; Bowman, J.P.; Fox, E.M. Novel Biocontrol Methods for
Listeria monocytogenes Biofilms in Food Production Facilities. Front. Microbiol. 2018, 9, 605. [CrossRef]
74. Borucki, M.K.; Peppin, J.D.; White, D.; Loge, F.; Call, D.R. Variation in Biofilm Formation among Strains of
Listeria monocytogenes. Appl. Environ. Microbiol. 2003, 69, 7336–7342. [CrossRef]
75. Yoon, Y.; Lee, H.; Lee, S.; Kim, S.; Choi, K.-H. Membrane fluidity-related adaptive response mechanisms of
foodborne bacterial pathogens under environmental stresses. Food Res. Int. 2015, 72, 25–36. [CrossRef]
76. Oxaran, V.; Dittmann, K.K.; Lee, S.H.I.; Chaul, L.T.; Fernandes de Oliveira, C.A.; Corassin, C.H.; Alves, V.F.;
Pereira De Martinis, E.C.; Gram, L. Behavior of Foodborne Pathogens Listeria monocytogenes and Staphylococcus
aureus in Mixed-Species Biofilms Exposed to Biocides. Appl. Environ. Microbiol. 2018, 84, e02038-18. [CrossRef]
77. Godreuil, S.; Galimand, M.; Gerbaud, G.; Jacquet, C.; Courvalin, P. Efflux pump Lde is associated
with fluoroquinolone resistance in Listeria monocytogenes. Antimicrob. Agents Chemother. 2003, 47, 704–708.
[CrossRef]
78. Jiang, X.; Yu, T.; Xu, P.; Xu, X.; Ji, S.; Gao, W.; Shi, L. Role of Efflux Pumps in the in vitro Development of
Ciprofloxacin Resistance in Listeria monocytogenes. Front. Microbiol. 2018, 9, 2350. [CrossRef]
79. Romanova, N.A.; Wolffs, P.F.G.; Brovko, L.Y.; Griffiths, M.W. Role of efflux pumps in adaptation and
resistance of Listeria monocytogenes to benzalkonium chloride. Appl. Environ. Microbiol. 2006, 72, 3498–3503.
[CrossRef]
80. Kovacevic, J.; Ziegler, J.; Wałecka-Zacharska, E.; Reimer, A.; Kitts, D.D.; Gilmour, M.W. Tolerance of
Listeria monocytogenes to Quaternary Ammonium Sanitizers Is Mediated by a Novel Efflux Pump Encoded by
emrE. Appl. Environ. Microbiol. 2016, 82, 939–953. [CrossRef]
81. Mata, M.T.; Baquero, F.; Pérez-Díaz, J.C. A multidrug efflux transporter in Listeria monocytogenes.
FEMS Microbiol. Lett. 2000, 187, 185–188. [CrossRef]
82. Quillin, S.J.; Schwartz, K.T.; Leber, J.H. The novel Listeria monocytogenes bile sensor BrtA controls expression
of the cholic acid efflux pump MdrT. Mol. Microbiol. 2011, 81, 129–142. [CrossRef] [PubMed]
83. Alcalde-Rico, M.; Hernando-Amado, S.; Blanco, P.; Martínez, J.L. Multidrug Efflux Pumps at the Crossroad
between Antibiotic Resistance and Bacterial Virulence. Front. Microbiol. 2016, 7, 1483. [CrossRef] [PubMed]
Pathogens 2020, 9, 528 11 of 12

84. Rodríguez-López, P.; Rodríguez-Herrera, J.J.; Vázquez-Sánchez, D.; Lopez Cabo, M. Current Knowledge on
Listeria monocytogenes Biofilms in Food-Related Environments: Incidence, Resistance to Biocides, Ecology
and Biocontrol. Foods 2018, 7, 85. [CrossRef] [PubMed]
85. Wales, D.A.; Davies, H.R. Co-Selection of Resistance to Antibiotics, Biocides and Heavy Metals, and Its
Relevance to Foodborne Pathogens. Antibiotics 2015, 4, 567–604. [CrossRef]
86. Romero, J.L.; Grande Burgos, M.J.; Pérez-Pulido, R.; Gálvez, A.; Lucas, R. Resistance to Antibiotics, Biocides,
Preservatives and Metals in Bacteria Isolated from Seafoods: Co-Selection of Strains Resistant or Tolerant to
Different Classes of Compounds. Front. Microbiol. 2017, 8, 1650. [CrossRef]
87. Bansal, M.; Nannapaneni, R.; Sharma, C.S.; Kiess, A. Listeria monocytogenes Response to Sublethal Chlorine
Induced Oxidative Stress on Homologous and Heterologous Stress Adaptation. Front. Microbiol. 2018, 9,
2050. [CrossRef]
88. Gravesen, A.; Jydegaard Axelsen, A.M.; Mendes da Silva, J.; Hansen, T.B.; Knøchel, S. Frequency of Bacteriocin
Resistance Development and Associated Fitness Costs in Listeria monocytogenes. Appl. Environ. Microbiol.
2002, 68, 756–764. [CrossRef]
89. NicAogáin, K.; O’Byrne, C.P. The Role of Stress and Stress Adaptations in Determining the Fate of the
Bacterial Pathogen Listeria monocytogenes in the Food Chain. Front. Microbiol. 2016, 7, 1865. [CrossRef]
90. Bucur, F.I.; Grigore-Gurgu, L.; Crauwels, P.; Riedel, C.U.; Nicolau, A.I. Resistance of Listeria monocytogenes to
Stress Conditions Encountered in Food and Food Processing Environments. Front. Microbiol. 2018, 9, 2700.
[CrossRef]
91. Lungu, B.; Group, A.; Milillo, S.R.; Johnson, M.G.; Crandall, P.G.; Ricke, S.C. Listeria monocytogenes: Antibiotic
Resistance in Food Production. Foodborne Pathog. Dis. 2011, 8, 569–578. [CrossRef]
92. Reygaert, W.C. An overview of the antimicrobial resistance mechanisms of bacteria. AIMS Microbiol. 2018, 4,
482–501. [CrossRef] [PubMed]
93. Smith, J.L.; Liu, Y.; Paoli, G.C. How does Listeria monocytogenes combat acid conditions? Can. J. Microbiol.
2012, 59, 141–152. [CrossRef] [PubMed]
94. Ryan, S.; Begley, M.; Gahan, C.; Hill, C. Molecular characterization of the arginine deiminase system in
Listeria monocytogenes: Regulation and role in acid tolerance. Environ. Microbiol. 2009, 11, 432–445. [CrossRef]
[PubMed]
95. Chorianopoulos, N.; Giaouris, E.; Grigoraki, I.; Skandamis, P.; Nychas, G.-J. Effect of acid tolerance response
(ATR) on attachment of Listeria monocytogenes Scott A to stainless steel under extended exposure to acid
or/and salt stress and resistance of sessile cells to subsequent strong acid challenge. Int. J. Food Microbiol.
2011, 145, 400–406. [CrossRef]
96. Karatzas, K.-A.G.; Suur, L.; O’Byrne, C.P. Characterization of the intracellular glutamate decarboxylase system:
Analysis of its function, transcription, and role in the acid resistance of various strains of Listeria monocytogenes.
Appl. Environ. Microbiol. 2012, 78, 3571–3579. [CrossRef] [PubMed]
97. Feehily, C.; Finnerty, A.; Casey, P.G.; Hill, C.; Gahan, C.G.M.; O’Byrne, C.P.; Karatzas, K.G. Divergent
Evolution of the Activity and Regulation of the Glutamate Decarboxylase Systems in Listeria monocytogenes
EGD-e and 10403S: Roles in Virulence and Acid Tolerance. PLoS ONE 2014, 9, e112649. [CrossRef]
98. Cotter, P.D.; Gahan, C.G.M.; Hill, C. Analysis of the role of the Listeria monocytogenes F0F1-ATPase operon in
the acid tolerance response. Int. J. Food Microbiol. 2000, 60, 137–146. [CrossRef]
99. Cotter, P.D.; Hill, C. Surviving the Acid Test: Responses of Gram-Positive Bacteria to Low pH. Microbiol. Mol.
Biol Rev. 2003, 67, 429–453. [CrossRef]
100. Shabala, L.; Lee, S.H.U.I.; Cannesson, P.; Ross, T.O.M. Acid and NaCl Limits to Growth of Listeria monocytogenes
and Influence of Sequence of Inimical Acid and NaCl Levels on Inactivation Kinetics. J. Food Prot. 2008, 71,
1169–1177. [CrossRef]
101. Sleator, R.D.; Gahan, C.G.M.; Hill, C. A Postgenomic Appraisal of Osmotolerance in Listeria monocytogenes.
Appl. Environ. Microbiol. 2003, 69, 1–9. [CrossRef]
102. Smiddy, M.; Sleator, R.D.; Patterson, M.F.; Hill, C.; Kelly, A.L. Role for Compatible Solutes Glycine Betaine
and l-Carnitine in Listerial Barotolerance. Appl. Environ. Microbiol. 2004, 70, 7555–7557. [CrossRef] [PubMed]
103. Chassaing, D.; Auvray, F. The lmo1078 gene encoding a putative UDP-glucose pyrophosphorylase is involved
in growth of Listeria monocytogenes at low temperature. FEMS Microbiol. Lett. 2007, 275, 31–37. [CrossRef]
[PubMed]
Pathogens 2020, 9, 528 12 of 12

104. Utratna, M.; Shaw, I.; Starr, E.; O’Byrne, C.P. Rapid, transient, and proportional activation of σ(B) in response
to osmotic stress in Listeria monocytogenes. Appl. Environ. Microbiol. 2011, 77, 7841–7845. [CrossRef] [PubMed]
105. Gardan, R.; Duché, O.; Leroy-Sétrin, S.; Labadie, J.; Consortium, E.L.G. Role of ctc from Listeria monocytogenes
in osmotolerance. Appl. Environ. Microbiol. 2003, 69, 154–161. [CrossRef]
106. Schmid, B.; Klumpp, J.; Raimann, E.; Loessner, M.J.; Stephan, R.; Tasara, T. Role of Cold Shock Proteins in
Growth of Listeria monocytogenes under Cold and Osmotic Stress Conditions. Appl. Environ. Microbiol. 2009,
75, 1621–1627. [CrossRef]
107. Eshwar, A.K.; Guldimann, C.; Oevermann, A.; Tasara, T. Cold-Shock Domain Family Proteins (Csps) Are
Involved in Regulation of Virulence, Cellular Aggregation, and Flagella-Based Motility in Listeria monocytogenes.
Front. Cell. Infect. Microbiol. 2017, 7, 453. [CrossRef]
108. Angelidis, A.S.; Smith, G.M. Role of the Glycine Betaine and Carnitine Transporters in Adaptation of Listeria
monocytogenes to Chill Stress in Defined Medium. Appl. Environ. Microbiol. 2003, 69, 7492–7498. [CrossRef]
109. Neunlist, M.R.; Federighi, M.; Laroche, M.; Sohier, D.; Delattre, G.; Jacquet, C.; Chihib, N. Cellular Lipid
Fatty Acid Pattern Heterogeneity between Reference and Recent Food Isolates of Listeria monocytogenes as a
Response to Cold Stress. Antonie Van Leeuwenhoek 2005, 88, 199–206. [CrossRef]
110. Beales, N. Adaptation of Microorganisms to Cold Temperatures, Weak Acid Preservatives, Low pH,
and Osmotic Stress: A Review. Compr. Rev. Food Sci. Food Saf. 2004, 3, 1–20. [CrossRef]
111. Lobo, V.; Patil, A.; Phatak, A.; Chandra, N. Free radicals, antioxidants and functional foods: Impact on
human health. Pharmacogn. Rev. 2010, 4, 118–126. [CrossRef]
112. Kocaman, N.; Sarimehmetoğlu, B. Stress responses of Listeria monocytogenes. Ankara Üniv. Vet. Fak. Derg.
2016, 63, 421–427.
113. Huang, Y.; Morvay, A.A.; Shi, X.; Suo, Y.; Shi, C.; Knøchel, S. Comparison of oxidative stress response and
biofilm formation of Listeria monocytogenes serotypes 4b and 1/2a. Food Control. 2018, 85, 416–422. [CrossRef]
114. Kayode, A.; Igbinosa, E.; Okoh, A.I. Overview of listeriosis in the Southern African Hemisphere-Review.
J. Food Saf. 2020, 40, e12732. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like